Hostname: page-component-848d4c4894-p2v8j Total loading time: 0.001 Render date: 2024-05-31T17:47:06.423Z Has data issue: false hasContentIssue false

Hypersonic shock impingement on a heated flat plate at Mach 7 flight enthalpy

Published online by Cambridge University Press:  07 December 2020

Eric Won Keun Chang
Affiliation:
Centre for Hypersonics, School of Mechanical and Mining Engineering, The University of Queensland, Brisbane, Queensland4072, Australia
Wilson Y. K. Chan
Affiliation:
Centre for Hypersonics, School of Mechanical and Mining Engineering, The University of Queensland, Brisbane, Queensland4072, Australia
Timothy J. McIntyre
Affiliation:
Centre for Hypersonics, School of Mathematics and Physics, The University of Queensland, Brisbane, Queensland4072, Australia
Ananthanarayanan Veeraragavan*
Affiliation:
Centre for Hypersonics, School of Mechanical and Mining Engineering, The University of Queensland, Brisbane, Queensland4072, Australia
*
Email address for correspondence: anandv@uq.edu.au

Abstract

Elevated wall temperatures and impinging shock interactions are prevalent features in hypersonic flight. Currently, there is a lack of literature regarding experimental studies examining both features in a flight-representative environment. This work details hot-wall, hypersonic, impinging shock/boundary-layer interaction experiments performed in the T4 Stalker Tube. The model configuration was a two-dimensional heated flat plate and a shock generator. The surface of the graphite flat plate was resistively heated to a mean temperature from $T_w=298\ \textrm {K}$ to $T_w\approx 675\ \textrm {K}$ during an experimental run. An oblique shock, generated by a plate that was inclined at $10^{\circ }$ or $12^{\circ }$ to the free stream, was impinged on the heated flat plate to induce boundary layer separation. The primary flow condition produced Mach 7 flight-equivalent nozzle-supply enthalpy with a unit Reynolds number of $4.93\times 10^6\ \textrm {m}^{-1}$. More flow conditions with lower unit Reynolds numbers and flow enthalpies were considered to examine flow separation characteristics. Schlieren and infrared thermography captured the flow field and the wall temperature distribution, respectively. The results showed that the size of the flow separation grew with a higher $T_w$ and a lower unit Reynolds number. Moreover, the scaled separation of the present data showed a high discrepancy with existing separation correlations developed from a supersonic impinging shock and a hypersonic compression ramp, mainly due to the higher shock strength. Instead, the present data followed a scaling law that includes the pressure ratio across the impinging shock with a slight dependence on the wall temperature ratio.

Type
JFM Rapids
Copyright
© The Author(s), 2020. Published by Cambridge University Press

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

REFERENCES

Billig, F. S. 1993 Research on supersonic combustion. J. Propul. Power 9 (4), 499514.CrossRefGoogle Scholar
Bleilebens, M. & Olivier, H. 2006 On the influence of elevated surface temperatures on hypersonic shock wave/boundary layer interaction at a heated ramp model. Shock Waves 15 (5), 301312.CrossRefGoogle Scholar
Boyce, R. R, Takahashi, M. & Stalker, R. J. 2005 a Mass spectrometric measurements of driver gas arrival in the T4 free-piston shock-tunnel. Shock Waves 14 (5–6), 371378.CrossRefGoogle Scholar
Boyce, R. R, Takahashi, M. & Stalker, R. J. 2005 b Mass spectrometric measurements of the freestream composition in the T4 free-piston shock-tunnel. Shock Waves 14 (5–6), 359370.CrossRefGoogle Scholar
Cebeci, T. & Bradshaw, P. 2012 Physical and Computational Aspects of Convective Heat Transfer. Springer.Google Scholar
Chan, W. Y. K., Jacobs, P. A., Smart, M. K., Grieve, S., Craddock, C. S. & Doherty, L. J. 2018 Aerodynamic design of nozzles with uniform outflow for hypervelocity ground-test facilities. J. Propul. Power 34 (6), 14671478.Google Scholar
Chang, E. W. K., Chan, W. Y. K., Hopkins, K. J., McIntyre, T. J. & Veeraragavan, A. 2020 Electrically-heated flat plate testing in a free-piston driven shock tunnel. Aerosp. Sci. Technol. 102, 111.Google Scholar
Chang, E. W. K., Chan, W. Y. K., McIntyre, T. J. & Veeraragavan, A. 2019 Hypersonic shock impingement on a heated surface in T4 free-piston driven shock tunnel. In 32nd International Symposium on Shock Waves, pp. 89–95. Research Publishing Services.Google Scholar
Currao, G. M. D., Choudhury, R., Gai, S. L., Neely, A. J. & Buttsworth, D. R. 2020 Hypersonic transitional shock-wave-boundary-layer interaction on a flat plate. AIAA J. 58 (2), 814829.CrossRefGoogle Scholar
Dolling, D. S. 2001 Fifty years of shock-wave/boundary-layer interaction research: what next? AIAA J. 39 (8), 15171531.Google Scholar
Elfstrom, G. M. 1972 Turbulent hypersonic flow at a wedge-compression corner. J. Fluid Mech. 53, 113127.CrossRefGoogle Scholar
Gai, S. L. & Khraibut, A. 2019 Hypersonic compression corner flow with large separated regions. J. Fluid Mech. 877, 471494.CrossRefGoogle Scholar
Holden, M. S. 1971 Establishment time of laminar separated flows. AIAA J. 9 (11), 22962298.CrossRefGoogle Scholar
Jacobs, P. A., Gollan, R. J., Potter, D. F., Zander, F., Gildfind, D. E., Blyton, P., Chan, W. Y. K. & Doherty, L. 2014 Estimation of high-enthalpy flow conditions for simple shock and expansion processes using the ESTCj program and library. Tech. Rep. 2011/02. Centre for Hypersonics, The University of Queensland.Google Scholar
Jaunet, V., Debieve, J. & Dupont, P. 2014 Length scales and time scales of a heated shock-wave/boundary-layer interaction. AIAA J. 52 (11), 2524–1741.CrossRefGoogle Scholar
Katzer, E. 1989 On the lengthscales of laminar shock/boundary-layer interaction. J. Fluid Mech. 206, 477496.CrossRefGoogle Scholar
Knisely, A. M. & Austin, J. M. 2016 Geometry and test-time effects on hypervelocity shock-boundary layer interaction. AIAA Paper 2016-1979.CrossRefGoogle Scholar
Mallinson, S. G., Gai, S. L. & Mudford, N. R. 1997 Establishment of steady separated flow over a compression–corner in a free–piston shock tunnel. Shock Waves 7 (4), 249253.CrossRefGoogle Scholar
Mee, D. J. 1993 Uncertainty analysis of conditions in the test section of the T4 shock tunnel. Tech. Rep. 1993/04. Centre for Hypersonics, University of Queensland.Google Scholar
Sandham, N. D., Schülein, E., Wagner, A., Willems, S. & Steelant, J. 2014 Transitional shock-wave/boundary-layer interactions in hypersonic flow. J. Fluid Mech. 752, 349382.CrossRefGoogle Scholar
Schülein, E. 2006 Skin friction and heat flux measurements in shock/boundary layer interaction flows. AIAA J. 44 (8), 17321741.CrossRefGoogle Scholar
Sriram, R. & Jagadeesh, G. 2015 Correlation for length of impinging shock-induced large separation bubble at hypersonic speed. AIAA J. 53 (9), 27712776.CrossRefGoogle Scholar
Sriram, R., Srinath, L., Devaraj, M. K. K. & Jagadeesh, G. 2016 On the length scales of hypersonic shock-induced large separation bubbles near leading edges. J. Fluid Mech. 806, 304355.CrossRefGoogle Scholar
Stalker, R. J., Paull, A., Mee, D. J., Morgan, R. G. & Jacobs, P. A. 2005 Scramjets and shock tunnels – the Queensland experience. Prog. Aerosp. Sci. 41 (6), 471513.CrossRefGoogle Scholar
Swantek, A. B. & Austin, J. M. 2015 Flowfield establishment in hypervelocity shock-wave/boundary-layer interactions. AIAA J. 53 (2), 311320.CrossRefGoogle Scholar
Volpiani, P. S., Bernardini, M. & Larsson, J. 2020 Effects of a nonadiabatic wall on hypersonic shock/boundary-layer interactions. Phys. Rev. Fluids 5, 120.CrossRefGoogle Scholar
Wagner, A., Schramm, J. M., Hannemann, K., Whitside, R. W. & Hickey, J. 2018 Hypersonic shock wave boundary layer interaction studies on a flat plate at elevated surface temperature. In Shock Wave Interactions (ed. Kontis, K.), pp. 231243.Google Scholar